• No results found

Meteorology For Coastal/Offshore Wind Energy In The United States: Recommendations And Research Needs For The Next 10 Years

N/A
N/A
Protected

Academic year: 2021

Share "Meteorology For Coastal/Offshore Wind Energy In The United States: Recommendations And Research Needs For The Next 10 Years"

Copied!
28
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

AMERICAN

METEOROLOGICAL

SOCIETY

Bulletin of the American Meteorological Society

EARLY ONLINE RELEASE

This is a preliminary PDF of the author-produced

manuscript that has been peer-reviewed and

accepted for publication. Since it is being posted

so soon after acceptance, it has not yet been

copyedited, formatted, or processed by AMS

Publications. This preliminary version of the

manuscript may be downloaded, distributed, and

cited, but please be aware that there will be

visual

differences

and possibly some content differences

between this version and the final published version.

The DOI for this manuscript is doi:

10.1175/BAMS-D-13-00108.1

The final published version of this manuscript will replace

the preliminary version at the above DOI once it is

available.

©

2013 American Meteorological Society

(2)

Meteorology For Coastal/Offshore Wind Energy In The United States:

1

Recommendations And Research Needs For The Next 10 Years

2

Cristina L. Archer (corresponding author) 3

University of Delaware 4

College of Earth, Ocean, and Environment

5 Newark, DE 19716 6 carcher@udel.edu 7 8 Brian A. Colle 9

Stony Brook University/SUNY, Stony Brook, New York 10

11

Luca Delle Monache 12

National Center for Atmospheric Research, Boulder, Colorado 13

14

Michael J. Dvorak 15

Sailor’s Energy, Berkeley, California 16

17

Julie Lundquist 18

University of Colorado at Boulder, and 19

National Renewable Energy Laboratory, Golden, Colorado 20

21

Bruce H. Bailey and Philippe Beaucage 22

AWS Truepower, LLC, Albany, New York 23

24

Matthew J. Churchfield 25

National Renewable Energy Laboratory, Golden, Colorado 26

27

Anna C. Fitch and Branko Kosovic 28

National Center for Atmospheric Research, Boulder, Colorado 29

30

Sang Lee and Patrick J. Moriarty 31

National Renewable Energy Laboratory, Golden, Colorado 32

33

Hugo Simao 34

Princeton University, Princeton, New Jersey 35

36

Richard J. A. M. Stevens 37

Johns Hopkins University, Baltimore, Maryland, and 38

University of Twente, Enschede (The Netherlands) 39

40

Dana Veron 41

University of Delaware, Newark, Delaware 42

43

John Zack 44

AWS Truepower, LLC, Albany, New York 45

(3)

Offshore wind energy is just starting in the United States, with imminent offshore wind 46

farms in Massachusetts, Maryland, and Rhode Island waters and with an ambitious goal 47

of 10 GW of installed offshore capacity by 2020 set by the U.S. Department of Energy 48

(DOE), which has recently funded seven “Advanced Technology Demonstration” 49

offshore wind projects to help achieve that goal. Although new in the U.S., offshore wind 50

energy began over 20 years ago in Europe and has now reached over 5.5 GW of installed 51

capacity worldwide, predominantly in Denmark and the United Kingdom. Given the 52

unfortunate coincidence of introducing a new industry during challenging economic 53

times, it is essential that public and private financial resources be effectively and 54

optimally directed towards those meteorological research needs that are emerging today 55

and that will be critical in the next decade. Identifying these research needs for wind 56

energy along the U.S. East Coast, both coastal and offshore, was the goal of a two-day 57

symposium held at the University of Delaware on 2728 February 2013. Over 40 58

participants gathered from academia, national laboratories, wind industry, and funding 59

agencies. 60

61

During the symposium, three main topics were explored: 1) wind resource assessment, 2) 62

wind power forecasting, and 3) turbulent wake losses. Overviews of the latest findings in 63

the three topics were given on the first day in the form of presentations, which were open 64

to students and the general public. On the second day, the experts gathered in a workshop 65

to identify research needs and provide recommendations for urgent action items. Whereas 66

specific research needs were identified for each of the three main topics, two emerged as 67

cross-cutting and urgent: 1) continuous, publicly available, multilevel measurements of 68

winds and temperature over U.S. offshore waters, and 2) quantification and reduction of 69

(4)

uncertainty. These two research needs and relevant recommendations (in italics) are 70

described first. 71

72

Research need #1: More offshore observations

73 74

Offshore meteorological measurements are challenging and expensive. Ideal 75

measurements would quantify the wind resource at several vertical levels spanning the 76

height of the turbine rotor disk to understand the rotor equivalent wind speed and possible 77

impacts on turbine power production. In European waters, designated research platforms 78

(e.g., FINO1 in Germany) have been established for characterization of offshore flow as 79

well as validation of new measurement technologies such as light detection and ranging 80

(lidar) and modeling approaches. The few long-term meteorological observations off the 81

East Coast are typically buoy-based, thereby restricting the altitude of wind 82

measurements to a few meters above the surface. A sparse network of nine towers, with 83

an elevation of ≤50 m, extends along the coast from Florida to Maine, but fails to provide 84

multilevel information and measurements at turbine hub-height or above. 85

Periodically, detailed measurements of wind and temperature have been conducted 86

offshore in short-term field campaigns, but the consistent long-term measurements 87

required for resource assessment are generally not available off the East Coast (with the 88

only exception being the Cape Wind tower in Nantucket Sound, Massachusetts). The 89

standard approach considered buoy measurements and then extrapolated them to higher 90

altitudes with assumptions of the shape of the wind profile (log-law or power-law). By 91

extrapolating surface or near-surface measurements with such smooth profiles, important 92

wind structures such as low-level jets are ignored. 93

(5)

The first recommendation is the deployment of a more dense network of

94

meteorological towers, which will enable traditional resource assessment measurements

95

such as wind speed, wind direction, and turbulence at several levels from the surface to 96

the rotor disk top, and temperature profiles for quantifying atmospheric stratification and 97

stability. Ideally, such towers could also provide a platform for validating remote sensing 98

measurements. The U.S. DOE has proposed the Reference Facility for Offshore 99

Renewable Energy (RFORE) to be located at the Chesapeake Light Tower, 100

approximately 13 miles off the Virginia Coast. The facility provides a first step towards 101

addressing the shortage of offshore wind data. 102

Beyond meteorological towers, remote sensing technology mounted either on fixed 103

towers or on floating platforms could provide data over broader regions. Scanning

104

Doppler lidar, wind-profiling lidar, and sodar can provide valuable wind speed and 105

direction measurements throughout the turbine rotor disk and beyond. Radiometers can 106

quantify temperature and humidity profiles to determine atmospheric stability. 107

In addition to long-term measurements of winds, temperature, and moisture profiles, 108

short-term intensive measurement campaigns with a broader deployment of instruments

109

would also be of value, especially for model validation.

110

These recommendations for more intensive observations extend a prior call for more 111

onshore meteorological observations and focused field campaigns made by DOE in 2008. 112

Since then, new types of remote sensing instruments have become more widely available 113

and more accepted in the wind energy industry for wind resource characterization. 114

115

Research need #2: Uncertainty characterization

116 117

(6)

Deterministic wind power forecasts based on numerical weather prediction (NWP) can 118

provide useful information for decision-making. However, by design, a single plausible 119

future state of the atmosphere starting from a single initial state is generated. Imperfect 120

initial and boundary conditions and model deficiencies inevitably lead to nonlinear error 121

growth during model integration. Accurate knowledge of the continuum of plausible 122

future states, the forecast probability density function (PDF), is considerably more useful 123

for decision-making because it allows for a quantification of the uncertainty associated 124

with a forecast. 125

“Ensembles” are used today to generate a set of plausible future atmospheric states 126

and to estimate the forecast PDF of atmospheric variables relevant to wind power. 127

Ensembles are created from the outputs of NWP models using any of the following: 128

various initial conditions, different parameterizations within a single model, stochastic 129

approaches with diverse numerical schemes, different models, and coupled ocean-130

atmosphere schemes. For wind energy, one important additional source of uncertainty 131

comes from the challenging step of wind-to-power conversion. 132

Ensembles are affected by biases in the ensemble mean and by lack of diversity 133

among the ensemble members, particularly in the planetary boundary layer (PBL). 134

Therefore, post-processing is an important component of the wind forecasting process 135

and should be explored further, preferably including methods and techniques developed

136

by the wind industry. Since the wind industry benefits from the findings published by the 137

research community and the public sector, it is recommended that a regular two-way 138

exchange of know-how between academia, public sector, and industry be established to

139

help advance the science and prevent the duplication of efforts. A promising post-140

processing technique is the analog approach, in which past observations that correspond 141

(7)

to past predictions that best match selected features of the current forecast, such as time 142

series of wind speed and direction, are used to correct the current forecast. Other 143

promising techniques are advanced model output statistics (e.g., neural networks, support 144

vector machines, and random forests). 145

Recently, operational centers have generated multiyear reforecast datasets to support 146

successful calibration of both deterministic and probabilistic forecasts. It is expected that 147

in the next few years new calibration techniques, possibly combining statistical and 148

dynamical approaches, will lead to large improvements in the accuracy of wind power 149

predictions and in the reliable characterization of their uncertainty. 150

151

Next, the three main topics and their specific research needs are described. 152

153

Topic #1: Resource assessment

154 155

Initial maps of the U.S. offshore wind resource from the National Renewable Energy 156

Laboratory (NREL) and others by Stanford University have identified gross 157

characteristics of the hub-height offshore wind resource, which have been generally 158

useful to policy makers and researchers and for early-stage project development. Using 159

mesoscale modeling techniques, these maps provide estimates of wind speed and 160

direction, diurnal and seasonal patterns, wind shear, and air density at horizontal grid 161

scales of approximately 1-5 km. This information, although essentially unverified due to 162

the lack of hub-height measurements described in Research Need #1, has enabled 163

numerous project siting studies, wind farm layout and energy production simulations, and 164

estimates of development potential as a factor of water depth, distance from shore, wind 165

(8)

resource, and other factors. However, there is a need to accurately capture dynamic 166

coastal processes, such as sea breezes, low-level jets, and other land-air-ocean

167

interactions, as they represent a significant source of variability in the available wind. 168

Data representing assessment periods of 2025 years (i.e., project lifetimes) are 169

typically required for bankable offshore projects; interannual speed variability of 4%6% 170

is not uncommon. The probability and magnitude of extreme events, particularly peak 171

winds and waves and hurricanes, and the effects of more common events, such as winter 172

storms, icing from sea spray, and salt corrosion, need to be better known to properly 173

design turbines and foundations and meet industry standards. In a changing climate, more 174

studies are needed to reduce the uncertainty of a changing wind resource as ocean,

175

offshore, and coastal temperatures change. Changes in the local wind environment over 176

time may also be caused by the increasing presence of other wind farms within a given 177

region, as described in Topic #3. 178

Recent studies have explored strategic temporal, climatological, and spatial aspects of 179

the offshore resource, including large-scale wind farm interconnection scenarios.U.S. 180

East Coast offshore wind has been found to be particularly coincident with peak-181

electricity demand. Similar studies should be performed to identify resource attributes 182

that can add value to generally higher offshore costs and evaluate the sensitivity of 183

project location, including distance from the shore, to load coincidence. 184

Significant offshore resource assessment uncertainties exist. Most of the 185

aforementioned studies relied on mesoscale modeling that was validated with generally 186

sparse in-situ data. Perhaps the largest uncertainty is extrapolating surface observations, 187

generally 5-m buoy anemometer measurements to heights across the turbine rotor. As 188

such, there is an urgent need for multilevel wind and temperature observations at 189

(9)

platforms offshore (as in Research Need #1), equipped with either meteorological towers

190

that are as tall or taller than hub height, or lidars. In the coastal region, transport 191

processes (advection of either maritime air inland or continental air offshore) during sea 192

and land breeze events often cause the PBL to deviate from classic well-mixed, neutrally-193

stable conditions. Existing PBL parameterizations struggle to perform well in these 194

conditions. Research effort is needed to improve such PBL parameterizations in coastal 195

regions.

196

Long-term wind climatologies require publicly available historic reanalysis data and 197

future climate data generated by models forced under different anthropogenic emission 198

scenarios. Most of the existing publicly available data are at a relatively coarse spatial 199

scale (>20 km) compared to the size of a typical wind farm. Dynamical downscaling 200

methods typically employ a regional climate model to generate higher spatio-temporal 201

wind climatologies but at a high computational expense for long climate records. 202

Stochastic downscaling methods are computationally cheaper and have been shown to 203

accurately downscale low-resolution reanalysis data with acceptable accuracy, as 204

compared to in-situ validation data. 205

206

Topic #2: Wind Power Forecasting

207 208

Wind power forecasting is challenging because the relationship between wind speed and 209

power production for a single wind turbine or a wind farm is nonlinear; for some wind 210

speed ranges, the sensitivity of power production forecasts to wind speed forecast error is 211

quite high. For example, a modest 1.5 m s-1 error in a wind speed forecast can, in some 212

cases, result in a power production forecast error of over 20% of a wind farm’s capacity. 213

(10)

A diverse set of prediction tools and input data have been applied to the wind power 214

forecast problem for a range of time scales. Intra-hour forecasts (060 minutes ahead) are 215

needed for regulation and real-time dispatch decisions. At this scale, the effects of small 216

eddies and turbulent mixing are important but cannot be resolved by operational models. 217

Therefore, mainly statistical methods are used, which are based on near real-time 218

observations. This has driven the deployment of meteorological sensors and lidars for 219

intra-hour forecasting. 220

The 1-6 hour-ahead forecast for load-following and next-operating hour commitment 221

has to account for various mesoscale weather phenomena (e.g., sea breezes, convective 222

systems, and local topography). The rapid-update NWP approach most likely offers the 223

best potential for improvement in this time frame. This is a tool with increasing 224

capability, largely because of improvements in data assimilation techniques (e.g., the 225

hybrid ensemble Kalman filter approach), the formulations of physics-based submodels, 226

and the amount and quality of data available for assimilation. The state of the art in rapid-227

update systems is the High-Resolution Rapid Refresh (HRRR) model, currently 228

undergoing experimental operation at the National Oceanic and Atmospheric 229

Administration, which assimilates the latest data and generates a 15-hour forecast on a 3-230

km grid every hour. 231

The day-ahead forecast is important for unit commitment, scheduling, and market 232

trading, which require knowledge of the evolving synoptic storm systems using NWP 233

models and ensembles. The seasonal predictions for resource planning and contingency 234

analysis require knowledge of global teleconnections (such as El Niño). These 235

predictions are based largely on the analysis of cyclical patterns and climate forecast 236

system models. 237

(11)

It is also recommended that more offshore observations be collected using towers,

238

lidars, and buoys, to better validate models, help with data assimilation and uncertainty

239

characterization, and improve the model physics, because many of the PBL schemes were 240

originally developed over land. These efforts will require a close collaboration between 241

operational forecast centers, industry, and academia. 242

Lastly, future efforts should focus on improving the models’ ability to represent the 243

PBL and the interactions of fine-scale processes with larger scale flows, both inland and

244

offshore. Such improvements will be possible only with investments that focus on 245

improving our understanding of these key processes using real observations. Several 246

workshops over the past twenty years have noted the need for improved PBL modeling, 247

but no concerted effort at making such improvements has been made. 248

249

Topic #3: Turbulent Wake Losses

250 251

Wind turbines generate wakes downstream, which are generally characterized by a wind 252

speed deficit and higher turbulence than the upwind environment. Because wakes can 253

reduce power production and increase structural fatigue in downstream turbines, 254

understanding wake properties, quantifying resulting power losses, and optimizing wind 255

turbine layouts to minimize such losses is especially important to the wind energy 256

industry. Accurately modeling turbine wakes is also important for other atmospheric 257

applications that span a wide range of spatial scales, such as the impacts of wind energy 258

deployment on the global climate, local meteorology, crop production, and the wind 259

resource itself. 260

(12)

Because atmospheric flows are characterized by high Reynolds numbers (~107-108), 261

the number of grid points required to explicitly resolve such flows with operating wind 262

turbines via direct numerical simulation is ~1018, which is prohibitive in the foreseeable 263

future. As such, the wind industry has traditionally relied on computationally efficient 264

wake models to simulate wind turbine wakes. In order of increasing complexities, these 265

earlier wake models include: analytical representations of the wake deficit (e.g., the 266

PARK model); parabolized forms of the Reynolds-Averaged Navier-Stokes (RANS) 267

equations (e.g., the Ainslie model, also called the eddy viscosity model; UPMPARK, 268

which uses a k-ɛ turbulence closure); hybrid models based on an internal boundary layer 269

growth parameterization and coupled with a parabolized RANS or an analytical model 270

(e.g., Deep-Array Wake Model and Large Array Wind Farm model); and nonlinear 271

RANS models (e.g., WindModeller, Ellipsys, and FUGA). Although these models are 272

attractive for their quick runtime, they have limited ability to capture the detailed wake 273

characteristics because they are not suitable for simulations of unsteady, anisotropic 274

turbulent flows. 275

To overcome these limitations, the research community has been using large-eddy 276

simulation (LES), in which large-scale flow structures are resolved while the effects of 277

smaller eddies are represented with a subgrid model (Smagorinsky or dynamical). In 278

addition, the wind turbine is represented by either an actuator disk (with or without 279

rotation features) or by actuator lines (one per blade) that exert a force on the flow and 280

act as a momentum sink, or by the vortex method. Arrays of multiple wind turbines, in 281

which multiple wakes interact with one another, have also been successfully simulated 282

with LES. However, because of the high CPU-hours required, LES can be conducted for 283

only a few hours or at equilibrium-state using periodic boundary conditions. 284

(13)

Because LES models for turbine wakes were traditionally developed in-house by 285

research centers or universities without any funds for distributing, maintaining, or testing 286

the codes, they are generally not available to the public. The only exception is the open-287

source Simulator for Offshore/Onshore Wind Farm Applications (SOWFA) from NREL, 288

which includes a finite-volume scheme, actuator disks/lines, and options for periodic or 289

nonperiodic boundary conditions. Although developing numerous in-house LES codes is 290

of value because researchers can obtain independent verification of results, it is 291

recommended that more effort and funds be devoted to maintaining LES codes for turbine

292

wakes and making them available to the public.

293

To avoid the steep computational costs of simulating real wind farms with high 294

numbers of turbines via LES, parameterizations of the effects of large wind farms on 295

regional meteorology and global climate have been developed for mesoscale NWP and 296

large-scale climate models, which are less computationally demanding. These 297

parameterizations represent wind farms as either an elevated momentum sink (often with 298

an added source of turbulent kinetic energy (TKE), increased surface roughness, or an 299

increased surface drag coefficient. Because surface-based parameterizations incorrectly 300

extract momentum near the surface, as opposed to around hub height, they are not

301

recommended for turbine wake impact studies. Although the global-scale impacts of even

302

high penetrations of wind energy have been proven negligible, local wakes extending 303

tens of kilometers downwind of individual large wind farms have been generated by 304

some wind farm parameterizations. However, to date, few observations are available to 305

verify these model results. 306

Comparing model results with wind tunnel experiments, with either a single turbine or

307

multiple turbines, is useful because the constant and controllable environment in a wind

(14)

tunnel can be reproduced well. However, wind tunnel conditions are different from real 309

atmospheric conditions and therefore field measurements are also recommended both at 310

individual turbines and at offshore wind farms. Short-term field campaigns, as well as

311

routine measurements (especially offshore) are needed to validate results under a large 312

umbrella of atmospheric conditions. It is recommended that inflow, near-wake, and far-313

wake vertical wind profiles and atmospheric stability be measured, as well as wake

314

properties, such as TKE and turbulent fluxes (preferably with scanning lidars or arrays of

315

sonic anemometers). 316

317

More information on offshore observations (Research need #1)

318 319

Cañadillas, B., A. Westerhellweg, and T. Neumann, 2011: Testing the performance of a 320

ground-based wind LiDAR system: One year intercomparison at the offshore platform 321

FINO1, DEWI Mag., 38, 58-64. 322

Chen, S. S., W. Zhao, M. A. Donelan, J. F. Price, and E. J. Walsh, 2007: The CBLAST-323

Hurricane Program and the Next-Generation Fully Coupled Atmosphere-Wave-Ocean 324

Models for Hurricane Research and Prediction. Bull. Amer. Meteor. Soc., 88, 311–317. 325

doi: http://dx.doi.org/10.1175/BAMS-88-3-311. 326

Clifton, A., S. Schreck, D. Jager, N. Kelley, and J. K. Lundquist, 2013: Meteorological 327

tower observations at the National Renewable Energy Laboratory. Journal of Solar 328

Energy Engineering. 135(3), 031017, doi:10.1115/1.4024068

329

Edson, J., and coauthors, 2007: The Coupled Boundary Layers and Air-Sea Transfer 330

Experiment in Low Winds. Bull. Amer. Meteor. Soc., 88, 341–356. doi: 331

http://dx.doi.org/10.1175/BAMS-88-3-341. 332

(15)

Friedrich, K. J., K. Lundquist, E. Kalina, M. Aitken, and R. Marshall, 2012: Stability and 333

Turbulence in the Atmospheric Boundary Layer: An Intercomparison of Remote 334

Sensing and Tower Observations. Geophys. Res. Lett., Vol. 39, No. 3, L03801, 335

doi:10.1029/2011GL050413. 336

Knupp, K., R. Ware, D. Cimini, F. Vandenberghe, J. Vivekanandan, E. Westwater, and 337

T. Coleman, 2009: Ground-Based Passive Microwave Profiling during Dynamic 338

Weather Conditions. J. Atmos. Oceanic Technol., 26, 1057-1073. 339

Muñoz-Esparza, D., B. Cañadillas, T. Neumann, and J. van Beeck, 2012: Turbulent 340

fluxes, stability and shear in the offshore environment: Mesoscale modelling and field 341

observations at FINO1. J. Renewable Sustainable Energy, 4, 063136, doi: 342

10.1063/1.4769201. 343

Neumann, T., K. Nolopp, and K. Herklotz, 2004: First operating experience with the 344

FINO1 research platform in the North Sea, DEWI Mag., 24, 27-34. 345

Nunalee, C. G., S. and Basu, 2013: Mesoscale modeling of coastal low-level jets: 346

implications for offshore wind resource estimation. Wind Energy, 347

doi: 10.1002/we.1628 348

Pichugina, Y. L., R. M. Banta, W. A. Brewer, S. P. Sandberg, and R. M. Hardesty, 2012: 349

Doppler Lidar-Based Wind-Profile Measurement System for Offshore Wind-Energy 350

and Other Marine Boundary Layer Applications. J. Appl. Meteor. Climatol., 51, 327– 351

349. 352

Schreck, S., J. K. Lundquist, W. Shaw, 2008: DOE Workshop Report: Research Needs 353

for Wind Resource Characterization. NREL Technical Report NREL/TP-500-43521. 354

(16)

Shaw, W. J., J. K. Lundquist, S. Schreck, 2009: Workshop on Research Needs for Wind 355

Resource Characterization. Bull. Amer. Meteorol. Soc., 90, 535-538, doi: 356

10.1175/2008BAMS2729.1 357

Wagner, R., I. Antoniou, S. M. Pedersen, M. S. Courtney, and H. E. Jørgensen, 2009: 358

The influence of the wind speed profile on wind turbine performance measurements. 359

Wind Energy, 12, 348–62.

360

Wharton, S., and J. K. Lundquist, 2012: Atmospheric stability affects wind turbine power 361

collection. Environ. Res. Lett., 7, 014005, doi:10.1088/1748-9326/7/1/014005. 362

363

More information on uncertainty characterization (Research need #2) and wind

364

power forecasting (Topic #2)

365 366

AMS, 2008: Enhancing weather information with probability forecasts: An information

367

statement of the American Meteorological Society. Bull. Amer. Meteor. Soc., 89,

368

1049–1053.

369

ANL, 2009: Wind power forecasting: State-of-the-art 2009. Decision and Information 370

Sciences Division Document ANL-DIS-10-1, 216 pp., available online at 371

www.dis.anl.gov/pubs/65613.pdf.

372

Berner, J., G. J. Shutts, M. Leutbecher, and T. N. Palmer, 2009: A spectral stochastic 373

kinetic energy backscatter scheme and its impact on flow-dependent predictability in 374

the ECMWF ensemble prediction system. J. Atmos. Sci., 66, 603–626. 375

Bishop, C. H., and Z. Toth, 1999: Ensemble transformation and adaptive observations. J. 376

Atmos. Sci., 56, 1748–1765.

(17)

Bowler, N. E., A. Arribas, K. R. Mylne, K. B. Robertson, and S. E. Beare, 2008: The

378

MOGREPS short-range ensemble prediction system. Quart. J. Roy. Meteor. Soc., 134,

379

703–722.

380

Buizza, R., M. J. Miller, and T. N. Palmer, 1999: Stochastic simulation of model 381

uncertainties in the ECMWF ensemble prediction system. Quart. J. Roy. Meteor. Soc., 382

125, 2887–2908.

383

Clifton, A., L. Kilcher, J. K. Lundquist, and P. Fleming, 2013: Using machine-learning to 384

predict wind turbine power output. Env. Res. Lett. 8 024009 doi:10.1088/1748-385

9326/8/2/024009. 386

Eckel, F. A., and C. F. Mass, 2005: Aspects of effective mesoscale, short-range ensemble 387

forecasting. Wea. Forecasting, 20, 328–350. 388

Erickson, M. E., B. A. Colle, and J. Charney, 2012: Impact of bias correction type and 389

conditional training on Bayesian model averaging over the northeast United States. 390

Wea. Forecasting, 27, 1449-1469.

391

European Wind Energy Association, 2013: The European offshore wind industry - Key 392

trends and statistics. European Wind Energy Association, Brussels, Belgium, 31 pp. 393

Hacker, J., S.-Y. Ha, C. Snyder, J. Berner, F. Eckel, E. Kuchera, M. Pocernich, S. Rugg, 394

J. Schramm, and X. Wang, 2011: The U.S. Air Force Weather Agency’s mesoscale 395

ensemble: Scientific description and performance results. Tellus A, 63, 625–641. 396

Hagedorn, R., T. M. Hamill, and J. S. Whitaker, 2008: Probabilistic forecast calibration

397

using ECMWF and GFS ensemble reforecasts. Part I: two-meter temperatures. Mon.

398

Wea. Rev., 136, 2608–2619.

(18)

Hamill, T. M., J. S. Whitaker, and X. Wei, 2004: Ensemble reforecasting: improving

400

medium-range forecast skill using retrospective forecasts. Mon. Wea. Rev., 132, 1434–

401

1447.

402

Hamill, T. M., and J. S. Whitaker, 2006: Probabilistic quantitative forecasts based on 403

reforecast analogs: theory and application. Mon. Wea. Rev., 134, 32093229.

404

Hamill, T. M., J. S. Whitaker, and S. L. Mullen, 2006: Reforecasting: improving skill 405

using retrospective forecasts. Bull. Amer. Meteor. Soc., 27, 33–46. 406

Hamill, T. M., and J. S. Whitaker, 2007: Ensemble calibration of 500-hPa geopotential

407

height and 850-hPa and 2-m temperatures using reforecasts. Mon. Wea. Rev., 135,

408

3273–3280.

409

Hamill, T. M., R. Hagedorn, and J. S. Whitaker, 2008: Probabilistic forecast calibration

410

using ECMWF and GFS ensemble forecasts. Part II: Precipitation. Mon. Wea. Rev.,

411

136, 2620–2632.

412

Hamill, T. M., G. T. Bates, J. S. Whitaker, D. R. Murray, M. Fiorino, T. J. Galarneau, Jr., 413

Y. Zhu, and W. Lapenta, 2012: NOAA’s second-generation global medium-range 414

ensemble 415

Hirschberg, P. A., E. Abrams, A. Bleistein, W. Bua, L. Delle Monache, T. W. Dulong, J.

416

E. Gaynor, B. Glahn, T. M. Hamill, J. A. Hansen, D. C. Hilderbrand, R. N. Hoffman, 417

B. H. Morrow, B. Philips, J. Sokich, and N. Stuart, 2011: A weather and climate 418

enterprise strategic implementation plan for generating and communicating forecast 419

uncertainty information. Bull. Amer. Meteor. Soc., 12, 1651–1666.

420

Holt, T., J. Pullen, and C. H. Bishop, 2009: Urban and ocean ensembles for improved 421

meteorological and dispersion modelling of the coastal zone. Tellus A, 61, 232–249. 422

(19)

Hou, D., E. Kalnay, and K. K. Droegemeier, 2001: Objective verification of the

423

SAMEX’98 ensemble forecasts. Mon. Wea. Rev., 129, 73–91.

424

Houtekamer, P. L., L. Lefaivre, J. Derome, H. Ritchie, and H. L. Mitchell, 1996: A 425

system simulation approach to ensemble prediction. Mon. Wea. Rev., 124, 1225–1242. 426

Houtekamer, P. L., H. L. Mitchell, G. Pellerin, M. Buehner, M. Charron, L. Spacek, and 427

B. Hansen, 2005: Atmospheric data assimilation with an ensemble Kalman filter: 428

Results with real observations. Mon. Wea. Rev., 133, 604–620. 429

Houtekamer, P. L., H. L. Mitchel, and X. Deng, 2009: Model error representation in an 430

operational Ensemble Kalman Filter. Mon. Wea. Rev., 137, 2126-2143. 431

Jones, M., B. A. Colle, and J. Tongue, 2007: Evaluation of a short-range ensemble 432

forecast system over the Northeast U.S., Wea. Forecasting, 22, 36-55. 433

Krishnamurti, T. N., C. M. Kishtawal, Z. Zhang, T Larow, D. Bachiochi, E. Williford, S. 434

Gadgil, and S. Surendran, 2000: Multimodel ensemble forecasts for weather and

435

seasonal climate. Journal of Climate, 13, 4196–4216.

436

Kuhl, D., I. Szunyogh, E. J. Kostelich, G. Gyarmati, D. J. Patil, M. Oczkowski, B. R. 437

Hunt, E. Kalnay, E. Ott, and J. A. Yorke, 2007: Assessing predictability with a local 438

ensemble Kalman filter. J. Atmos. Sci., 64, 1116–1140. 439

Lorenz, E. N., 1963: Deterministic non-periodic flow. J. Atmos. Sci., 20, 130–141.

440

Marquis, M., J. Wilczak, M. Ahlstrom, J. Sharp, A. Stern, J. C. Smith, and S. Calvert, 441

2011: Forecasting the wind to reach significant penetration levels of wind energy. 442

Bull. Amer. Meteor. Soc., 92, 1159–1171. 443

Molteni, F., R. Buizza, T. N. Palmer, and T. Petroliagis, 1996: The new ECMWF

444

ensemble prediction system: Methodology and validation. Quart. J. Roy. Meteor. Soc.,

445

122, 73–119.

(20)

NRC, 2006: Completing the Forecast: Characterizing and Communicating Uncertainty

447

for Better Decisions Using Weather and Climate Forecasts. National Academies

448

Press, 124 pp.

449

NREL, 2011: Wind energy forecasting: A collaboration of the National Center for 450

Atmospheric Research (NCAR) and Xcel Energy, available online at 451

http://www.nrel.gov/docs/fy12osti/52233.pdf. 452

Perlin, N., E.D. Skyllingstad, R.M. Samelson, and P.L. Barbour, 2007: Numerical 453

simulation of air–sea coupling during coastal upwelling. Journal of Physical 454

Oceanography, 37, 2081–2093.

455

Stensrud, D. J., J. W. Bao, and T. T. Warner, 2000: Using initial condition and model 456

physics perturbations in short-range ensemble simulations of mesoscale convective 457

systems. Mon. Wea. Rev., 128, 2077–2107. 458

Teixeira, J., and C. A. Reynolds, 2008: Stochastic nature of physical parameterizations in 459

ensemble prediction: A stochastic convection approach. Mon. Wea. Rev., 136, 483– 460

496. 461

Thomas, S. J., J. P. Hacker, M. Desgagné, and R. Stull, 2002: An ensemble analysis of

462

forecast errors related to floating point performance. Wea. Forecasting, 17, 898–906.

463

Toth, Z., and E. Kalnay, 1993: Ensemble forecasting at NMC: The generation of

464

perturbations. Bull. Am. Meteorol. Soc., 74, 2317–2330.

465

Toth, Z., and E. Kalnay, 1997: Ensemble forecasting at NCEP: The breeding method.

466

Mon. Wea. Rev., 125, 3297–3319.

467

Wandishin, M. S., S. L. Mullen, D. J. Stensrud, and H. E. Brooks, 2001: Evaluation of a

468

short-range multi-model ensemble system. Mon. Wea. Rev., 129, 729–747.

(21)

WMO, 2008: Guidelines on communicating forecast uncertainty. WMO Tech. Document

470

4122, 25 pp, available online at www.wmo.int/pages/prog/amp

/pwsp/documents/TD-471

1422.pdf.

472

Wilks, D. S., and T. M. Hamill, 2007: Comparison of ensemble-MOS methods using

473

GFS reforecasts. Mon. Wea. Rev., 135, 2379–2390.

474

Wilks, D. S., 2009: Extended logistic regression to provide full-probability-distribution

475

MOS forecasts. Meteorol. Appl., 16, 361–368.

476 477

More information on resource assessment (Topic #1)

478 479

Brower, M., 2012: Wind resource assessment: a practical guide to developing a wind 480

project. John Wiley & Sons, Hoboken (New Jersey), 280 pp. 481

Delle Monache, L., T. Nipen, Y. Liu, G. Roux, and R. Stull, 2011: Kalman filter and 482

analog schemes to post-process numerical weather predictions. Mon. Wea. Rev., 139, 483

3554-3570. 484

Dvorak, M. J., E. D. Stoutenburg, C. L. Archer, W. Kempton, and M. Z. Jacobson, 2012: 485

Where is the ideal location for a US East Coast offshore grid? Geophysical Research 486

Letters, 39, L06804, 6 pp., doi:10.1029/2011GL050659.

487

Dvorak, M. J., B. A. Corcoran, J. E. Ten Hoeve, N. G. McIntyre, M. Z. Jacobson, 2013: 488

US East Coast offshore wind energy resources and their relationship to peak-time 489

electricity demand. Wind Energy, in press. 490

Kempton, W., C. L. Archer, A. Dhanju, R. W. Garvine, and M. Z. Jacobson, 2007: Large 491

CO2 reductions via offshore wind power matched to inherent storage in energy

end-492

uses. Geophysical Research Letters, 34, L02817, doi:10.1029/2006GL028016. 493

(22)

Kempton, W., F. Pimenta, D. E. Veron, and B. A. Colle, 2010: Electric power from 494

offshore wind via synoptic-scale interconnection. Proc. Natl. Acad. Sci., 107,7240. 495

496

More information on wake characteristics and losses (Topic #3)

497 498

Adams, A. S., and D. W. Keith, 2013: Are global wind power resource estimates 499

overstated? Environ. Res. Lett. 8, 015021. 500

Ainslie, J. F., 1988: Calculating the flowfield in the wake of wind turbines. J. 501

Wind Eng. Ind. Aerod., 27, 213-224.

502

American Meteorological Society Offshore Wind Energy Committee, 2013: The 503

need for expanded meteorological and oceanographic data to support resource 504

characterization and design condition definition for offshore wind power 505

projects in the United States. Available at: 506

http://www.ametsoc.org/boardpges/cwce/docs/BEP/Offshore-Wind/2013-05-507

Offshore-Wind-APT-Final-Report.pdf 508

Baidya Roy, S., S. W. Pacala, and R. L. Walko, 2004: Can large wind farms affect 509

local meteorology? J. Geophys. Res., 109, D19101. 510

Baidya Roy S, 2011: Simulating impacts of wind farms on local 511

hydrometeorology. J. Wind Eng. Ind. Aerodyn. doi:10.1016. 512

Barrie D., and D. B. Kirk-Davidoff, 2010: Weather response to a large wind 513

turbine array. Atmos. Chem. Phys., 10, 769-775. 514

Barthelmie, R. J., G. C. Larsen, S. T. Frandsen, L. Folkerts, K. Rados, S. C. Pryor, 515

B. Lange, G. Schepers, 2006: Comparison of Wake Model Simulations with 516

(23)

Offshore Wind Turbine Wake Profiles Measured by Sodar. J. Atmos. Oceanic 517

Technol., 23, 888–901.

518

Barthelmie, R. J., K. Hansen, S.T. Frandsen, O. Rathmann, J. G. Schepers, W. 519

Schlez, J. Phillips, K. Rados, A. Zervos, E.S. Politis, and P. Chaviaropoulos, 520

2009: Modelling and Measuring Flow and Wind Turbine Wakes in Large 521

Wind Farms Offshore, Wind Energy 12, 431. 522

Beaucage, P., N. Robinson, M. Brower, and C. Alonge, 2012: Overview of six 523

commercial and research wake models for large offshore wind farms. 524

Proceeding, EWEA conference, Copenhagen. 525

Cal, R. B., J. Lebrón, L. Castillo, H. S. Kang, and C. Meneveau, 2010: 526

Experimental study of the horizontally averaged flow structure in a model 527

wind-turbine array boundary layer. J. Renewable Sustainable Energy, 2, 528

013106. 529

Calaf, M., C. Meneveau, and J. Meyers, 2010: Large eddy simulation study of 530

fully developed wind-turbine array boundary layers. Phys. Fluids, 22, 015110. 531

Calaf, M., M. B. Parlange, and C. Meneveau, 2011: Large eddy simulation study 532

of scalar transport in fully developed wind-turbine array boundary layers. Phys. 533

Fluids, 23, 126603.

534

Chamorro, L.P., and F. ort -Agel, 2009: A tunnel investigation of wind-535

turbine wakes: boundary-layer turbulence effects. Bound.-Lay. Meteor., 132, 536

129–149. 537

Churchfield, M. J., S. Lee, S., J. Michalakes, and P. Moriarty, 2012: A numerical 538

study of the effects of atmospheric and wake turbulence on wind turbine 539

dynamics. J. Turbulence, 13(14), 1-32. 540

(24)

Churchfield, M. J., S. Lee, P. Moriarty, L. A. Martinez, S. Leonardi, G. 541

Vijayakumar, and J. G. Brasseur, 2012: A large-eddy simulation of wind-plant 542

aerodynamics, 50th AIAA Aerospace Sciences Meeting including the New 543

Horizons Forum and Aerospace Exposition, Nashville, Tennessee, 912 544

January 2012. 545

Crespo, A., L. Chacón, J. Hernández, F. Manuel, and J. Grau, 1994: UPMPARK: 546

A Parabolic 3D Code to Model Wind Farms, Proceeding, European Wind 547

Energy Conference 1994, Thessalonica, Greece, 10-14 October 1994,

454-548

459. 549

Fitch, A. C., J. B. Olson, J. K. Lundquist, J. Dudhia, A. K. Gupta, J. Michalakes, 550

and I. Barstad, 2012: Local and mesoscale impacts of wind farms as 551

parameterized in a mesoscale NWP model. Mon. Weather Rev., 140, 3017-552

3038. 553

Fitch, A. C., J. B. Olson, and J. K. Lundquist, 2013: Mesoscale influences of wind 554

farms throughout a diurnal cycle. Mon. Weather Rev., 141, 2173-2198. 555

http://dx.doi.org/10.1175/MWR-D-12-00185.1. 556

Fitch, A. C., J. B. Olson, and J. K. Lundquist, 2013: Parameterization of wind 557

farms in climate models. J. Climate. http://dx.doi.org/10.1175/JCLI-D-12-558

00376.1. 559

Ivanell, S. S. A., 2009: Numerical computations of wind turbine wakes, PhD 560

Thesis, Gotland University, Stockholm, Sweden. 561

Ivanell, S. S. A., R. Mikklesen, J. Sorensen, and D. Henningson, 2008: Validation 562

of methods using EllipSys3D. Technical report, KTH, TRITA-MEK 2008:12. 563

(25)

Garza, J., A. Blatt, R. Gandoin, and S.-Y. Hui, 2011: Evaluation of two novel 564

wake models in offshore wind farms. Proceeding, EWEA Offshore 2011, 565

Amsterdam, The Netherlands, 29 November1 December 2011. 566

Jacobson, M. Z., and C. L. Archer, 2012: Saturation wind power potential and its 567

implications for wind energy. Proc. Natl. Acad. Sci., 109(39), 15681. 568

Jensen, N. O., 1983: A Note on Wind Generator Interaction, Risø National 569

Laboratory, Risø-M-2411 570

Jimenez, A., A. Crespo, E. Migoya, and J. Garcia, 2007: Advances in large-eddy 571

simulation of a wind turbine wake. J. Phys.: Conf. Ser., 75, 012041. 572

Jimenez, A., A. Crespo, E. Migoya, and J. Garcia, J., 2008: Large-eddy simulation 573

of spectral coherence in a wind turbine wake, Environ. Res. 3, 015004. 574

Jimenez, A., A. Crespo, and E. Migoya, 2010: Application of a LES technique to 575

characterize the wake deflection of a wind turbine in yaw, Wind Energy, 13, 576

559-572. 577

Katic, I., J. Højstrup, and N. O. Jensen, 1987: A simple model for cluster 578

efficiency, proceeding, European Wind Energy Conference 1986, Rome, 1, 579

407-410. 580

Keith, D. W., J. F. DeCarolis, D. C. Denkenberger, D. H. Lenschow, S. L. 581

Malyshev, S. Pacala, and P. J. Rasch, 2004: The influence of large-scale wind 582

power on global climate. Proc. Natl. Acad. Sci., 101(46), 16115–16120. 583

Kirk-Davidoff, D. B., and D. W. Keith, 2008: On the climate impact of surface 584

roughness anomalies. J. Atmos. Sci. 85, 2215-2234. 585

Kosovic, B., and J. D. Mirocha, 2012: “Toward High-resolution Multiscale 586

Simulations of Flows in Heterogeneous Atmospheric Boundary Layers.” 20th 587

(26)

Symposium on Boundary Layers and Turbulence, Boston, MA, 9-13 July, 588

2012. 589

Lee, S., Churchfield, M., P. Moriarty, J. Jonkman, and J. Michalakes, 2013: A 590

numerical study of atmospheric and wake turbulence impacts on wind turbine 591

fatigue loadings. J Solar Energy Engineering, 135(3), 031001. 592

Madsen, H. Aa., G. C. Larsen, T. J. Larsen, N. Troldborg, and R. Mikkelsen, 593

2010: Calibration and validation of the dynamic wake meandering model 594

implemented in the aeroelastic code HAWC2. J. Solar Energy Engineering, 595

132(4), 041014.

596

Mann, J., 1994: The spatial structure of neutral atmospheric surface-layer 597

turbulence. J. Fluid Mech., 273, 141–168. 598

Martinez, L. A., S. Leonardi, M. Churchfield, and P. Moriarty, 2012: A 599

comparison of actuator disk and actuator line wind turbine models and best 600

practices for their use, 50th AIAA Aerospace Sciences Meeting including the 601

New Horizons Forum and Aerospace Exposition, Nashville, Tennessee, 912 602

January 2012. 603

Marvel K., B. Kravitz, and K. Caldeira, 2012: Geophysical limits to global wind 604

power. Nature Climate Change, 3, 118–121. 605

Meyers J., and C. Meneveau, 2011: Optimal turbine spacing in fully developed 606

wind farm boundary layers. Wind Energy, 15, 305-317. 607

Mikkelsen, R., 2003: Actuator disk methods applied to wind turbines. Ph. D. 608

Thesis, Technical University of Denmark. 609

(27)

Mikkelsen, R., J. N. Sørensen, J. A. Michelsen, and N. Troldborg, 2005: 610

Aeroelastic actuator line modeling, proceeding, 18th IEA Symposium on 611

Aerodynamics of Wind Turbines, Pamplona, Spain, 2005. 612

National Renewable Energy Laboratory, “SOWFA: Simulator for On/Offshore 613

Wind Farm Applications,” 614

http://wind.nrel.gov/designcodes/simulators/sowfa/. 615

Rajewski, D., and coauthors, 2013: Crop Wind Energy Experiment (CWEX): 616

Observations of Surface-Layer, Boundary Layer, and Mesoscale Interactions 617

with a Wind Farm. Bull Amer Meteor Soc 94: 655–672. 618

Rhodes, M. E., and J. K. Lundquist, 2013: The Effect of Wind Turbine Wakes on 619

Summertime Midwest Atmospheric Wind Profiles. In press at Boundary-620

Layer Meteorology. doi:10.1007/s10546-013-9834-x

621

Sanderse, B., S. P. van der Pijl, and B. Koren, 2011: Review of computational 622

fluid dynamics for wind turbine wake aerodynamics. Wind Energy, 14, 799-623

819. 624

Sørensen, N. N., 2003: General Purpose Flow Solver Applied to Flow over Hills, 625

Ph.D. Thesis, Risø-R-827(EN), Risø National Laboratory, Roskilde, Denmark. 626

Sørensen, J. N., and W. Z. Shen, 2002: Numerical modeling of wind turbine 627

wakes. J. Fluids Engineering, 124(2), 393-399. 628

Thomsen, K., 2001: Aeroelastic load simulations, FLEX5 sample, Abstract from 629

C-WET expert training RISO, Risø, Denmark. 630

Troldborg, N., G. C. Larsen, H. A. Madsen, K. S. Hansen, J. N. Sørensen, and R. 631

Mikkelsen, 2010: Numerical simulations of wake interactions between two 632

wind turbines at various inflow conditions. Wind Energy, 14, 859-876. 633

(28)

Vermeer, L. J., J. N. Sørensen, and A. Crespo, A., 2003: Wind turbine wake 634

aerodynamics. Prog. Aerosp. Sci., 39, 467–510. 635

Wagner, R., M. Courtney, J. Gottschall, and P. Lindellow-Marsden, 2011; 636

Accounting for the speed shear in wind turbine power performance 637

measurement. Wind Energy, 14(8), 993-1004, doi:10.1002/we.509. 638

Wang, C., and R. J. Prinn, 2010: Potential climatic impacts and reliability of very 639

large-scale wind farms. Atmos. Chem. Phys. 10, 2053-2061. 640

Wu, Y.-T., and F. Porté-Agel, 2011: Large-eddy simulation of wind-turbine 641

wakes: evaluation of turbine parameterizations. Bound.-Lay. Meteor., 138, 642

345–366. 643

Wu, Y.-T., and F. Porté-Agel, 2013: Simulation of turbulent flow inside and 644

above wind farms: model validation and layout effects. Bound.-Lay. Meteor., 645

146, 181-205.

646

Yang, X., S. Kang, and F. Sotiropoulos F., 2012: Computational study and 647

modeling of turbine spacing effects in infinite aligned wind farms. Phys. 648

Fluids, 24, 115107.

649

Zhou, L., Y. Tian, S. Baidya Roy, C. Thorncroft, L. F. Bosart, and Y. Hu, 2012: 650

Impacts of wind farms on land surface temperature. Nature Climate Change, 2, 651

539–543, doi:10.1038. 652

Referenties

GERELATEERDE DOCUMENTEN

The large-scale ocean circulation system is expected to weaken in a warmer climate (Gregory et al., 2005 ) and will lead to higher sea levels in the northern North Atlantic

The main question of this thesis was answered by investigating the methods collegia and trading communities used to increase their visibility and the specific locations

This thesis focusses on crew transfer vessels that service the offshore wind turbines and the technicians required for each maintenance task.. Industry professionals are interviewed

Since the average costs of offshore wind power are high and it is not possible to make profits without subsidies, it is necessary to discuss the valuation according financial

Turbine foundations were mainly used for roosting, but occasionally, great black-backed and herring gulls were observed foraging on mussels growing on the lower reaches of the

He/she composes the folktale, performs it orally and assur·es that the folktale is transmitted fr·om individual to individual; from generation to generation; and

Dit is teen die agtergrond van hierdie ontluikende nuwe wêreld met ’n groter premie op innoverende aanpassing, verskeidenheid en die noodsaak vir die ontwikkeling van

Chapter 8 by Jakkie Cilliers (Stability and security in Southern Africa) provides and overview of the state of the nation in Southern Africa, as well as an update and cursory